+ All Categories
Home > Documents > Detection of Changes in Multivariate Time Series With...

Detection of Changes in Multivariate Time Series With...

Date post: 30-May-2020
Category:
Upload: others
View: 6 times
Download: 0 times
Share this document with a friend
29
Detection of Changes in Multivariate Time Series With Application to EEG data * Claudia Kirch , Birte Muhsal , Hernando Ombao § . December 19, 2013 Abstract The primary contributions of this paper are rigorously developed novel statistical methods for detecting change points in multivariate time series. We extend the class of score type change point statistics considered by [Huˇ skov´ a et al., 2007] to the vector autoregressive (VAR) case and the epidemic change alternative. Our proposed procedures do not require the observed time series to actually follow the VAR model. Instead, following the strategy implicitly employed by practitioners, our approach takes model misspecification into account so that our detection procedure uses the model background merely for feature extraction. We derive the asymptotic distributions of our test statistics and show that our procedure has asymptotic power of 1. The proposed test statistics require the estimation of the inverse of the long-run covariance matrix which is particularly difficult in higher-dimensional settings (i.e., where the dimension of the time series and the dimension of the parameter vector are both large). Thus we robustify the proposed test statistics and investigate their finite sample properties via extensive numerical experiments. Finally, we apply our procedure to electroencephalograms and demonstrate its potential impact in identifying change points in complex brain processes during a cognitive motor task. Keywords: Change points; multivariate time series; epidemic change; vector autoregressive model; EEG data; AMS Subject Classification 2010: 1 Introduction The primary contributions of this paper are rigorously developed novel statistical methods for detecting change points in multivariate time series. Change point detection is crucial in analyzing many time series * This work was supported by DFG grant KI 1443/2-2, US National Science Foundation (Division of Mathematical Sciences) and the Stifterverband f¨ ur die Deutsche Wissenschaft by funds of the Claussen-Simon-trust which also financed the position of the first author. Karlsruhe Institute of Technology (KIT), Institute of Stochastics, Kaiserstr. 89, 76133 Karlsruhe, Germany; [email protected] Karlsruhe Institute of Technology (KIT), Institute of Stochastics, Kaiserstr. 89, 76133 Karlsruhe, Germany; [email protected] § University of California at Irvine, Department of Statistics, Irvine CA 92697, USA; [email protected] 1
Transcript
Page 1: Detection of Changes in Multivariate Time Series With ...ckirch/seite/publications/media/pp_kmo_1… · Detection of Changes in Multivariate Time Series With Application to EEG data

Detection of Changes in Multivariate Time Series With Application to

EEG data ∗

Claudia Kirch†, Birte Muhsal‡, Hernando Ombao§.

December 19, 2013

Abstract

The primary contributions of this paper are rigorously developed novel statistical methods for

detecting change points in multivariate time series. We extend the class of score type change point

statistics considered by [Huskova et al., 2007] to the vector autoregressive (VAR) case and the epidemic

change alternative. Our proposed procedures do not require the observed time series to actually follow

the VAR model. Instead, following the strategy implicitly employed by practitioners, our approach

takes model misspecification into account so that our detection procedure uses the model background

merely for feature extraction. We derive the asymptotic distributions of our test statistics and show

that our procedure has asymptotic power of 1. The proposed test statistics require the estimation of the

inverse of the long-run covariance matrix which is particularly difficult in higher-dimensional settings

(i.e., where the dimension of the time series and the dimension of the parameter vector are both large).

Thus we robustify the proposed test statistics and investigate their finite sample properties via extensive

numerical experiments. Finally, we apply our procedure to electroencephalograms and demonstrate its

potential impact in identifying change points in complex brain processes during a cognitive motor task.

Keywords: Change points; multivariate time series; epidemic change; vector autoregressive model;

EEG data;

AMS Subject Classification 2010:

1 Introduction

The primary contributions of this paper are rigorously developed novel statistical methods for detecting

change points in multivariate time series. Change point detection is crucial in analyzing many time series

∗This work was supported by DFG grant KI 1443/2-2, US National Science Foundation (Division of Mathematical

Sciences) and the Stifterverband fur die Deutsche Wissenschaft by funds of the Claussen-Simon-trust which also financed

the position of the first author.†Karlsruhe Institute of Technology (KIT), Institute of Stochastics, Kaiserstr. 89, 76133 Karlsruhe, Germany;

[email protected]‡Karlsruhe Institute of Technology (KIT), Institute of Stochastics, Kaiserstr. 89, 76133 Karlsruhe, Germany;

[email protected]§University of California at Irvine, Department of Statistics, Irvine CA 92697, USA; [email protected]

1

Page 2: Detection of Changes in Multivariate Time Series With ...ckirch/seite/publications/media/pp_kmo_1… · Detection of Changes in Multivariate Time Series With Application to EEG data

data – ignoring them leads to incorrect conclusions and forecasts. Thus, experts in finance keenly watch

out for changes in several market indicators because these could indicate impact of government policy

or mark a need to revise long term investment strategies. Moreover, neuroscientists study the effect of

external stimuli on neuronal responses by tracking changes in brain signals such as electroencephalograms.

Brain signals are highly dynamic since these are realizations of complex cognitive processes. In this context

detection of change points is necessary in order to understand how a cognitive process unfolds in response

to a stimulus.

Consider two background processes, denoted Π1 and Π2, that run in parallel but for each time point

only one of these two is activated. In other words, when Π1 is activated at time t then Π2 is turned

off. These two processes are identifiable in the sense that they have different second order structure and

thus our goal is to search for the point(s) in time when the covariance structure changes. Here, we shall

investigate two scenarios: the at-most-one-change (AMOC) and the epidemic change. In the AMOC

situation, the goal is to estimate the time t1 when Π1 is turned off (equivalently, the time when Π2 is

activated). In the epidemic change situation, change point t1 corresponds to the time when Π1 is turned

off (equivalently, when Π2 is activated) and change point t2 corresponds to the time Π2 is turned off

thus returning to Π1. Here, the goal is to estimate the two change points t1 and t2. To identify change

points, or to analyze time series data in general, there are many possible stochastic representations or

time series models that could be utilized. Based on both theoretical and practical considerations, we shall

develop methods that use autoregressive (AR) models. As studied by [Bickel and Buhlman, 1999] in the

context of bootstrap methods, linear stationary processes can be well approximated by autoregressive

(AR) processes that have sufficiently large order. Moreover, the AR model has been utilized to analyze

many real-life time series data such as electroencephalograms ([Pfurtscheller and Haring, 1972]) seismic

signals ([Lesage et al., 2002]) and speech signals ([Vermaak et al., 2002]). While AR models do not fully

capture the features of many complex time series data, here, we do not claim that the observed time

series data is a realization of some AR model. Rather, the AR model will be utilized as a tool for feature

extraction whose parameters capture changes in the second order structure of the time series.

We provide the theory for a large class of statistics, which will be of interest in many different areas.

However, in practice it is of utter importance to adapt the test statistic to the actual data being analyzed.

We will demonstrate how to do this throughout the paper using the data example that will be examined in

detail in Section 4. Our data consists of electroencephalograms (EEG) recorded from a single participant

in a visual-motor task experiment. In this experiment, the participant was instructed to move the hand-

held joystick either towards the left of center or towards the right of center. Each trial consists of one

second (total of n = 512 time points) whose midpoint is approximately the point in time when the stimulus

was presented (at t = 250). Thus, each trial consists of half a second of EEG recording before stimulus

presentation and another half a second post stimulus presentation. Change points are expected on the

post-stimulus recording as the brain integrates and processes information. Indeed the proposed method

clearly captures these changes in brain dynamics. We note that our EEG data example serves only as an

illustration – the methodology has broad applicability and is by no means limited to EEG data.

To describe our proposed procedure, we first give a brief introduction on the vector autoregressive

model. Let Yt = [Yt(1), . . . , Yt(d)]> be a d−dimensional vector recorded at time t across all d channels.

Then Yt is said to follow a vector autoregressive model of order p, denoted VAR(p), if it can be expressed

as Yt = α1Yt−1 + . . . + αpYt−p + εt where α`, ` = 1, . . . , p, are the d× d autoregressive matrices

2

Page 3: Detection of Changes in Multivariate Time Series With ...ckirch/seite/publications/media/pp_kmo_1… · Detection of Changes in Multivariate Time Series With Application to EEG data

corresponding to the lagged values of Yt; and εt is white noise. An equivalent formulation of the VAR(p)

model is given as follows. Define

Yt−1 = [Yt−1(1)>, . . . ,Yt−1(d)>]> where Yt−1(j) = [Yt−1(j), . . . , Yt−p(j)]>, j = 1, . . . , d.

Here, dim(Yt−1) = pd × 1 and dim(Yt−1(i)) = p × 1. Next define the matrix Φ =

Φ>(1)

. . .

Φ>(d)

with

dim(Φ) = d× dp whose i-th row is given by

Φ>(i) = [φ1,1(i), . . . , φ1,p(i), . . . , φd,1(i), . . . , φd,p(i)].

The conditional mean of the i-th channel Yt(i) is a linear combination of values at all channels from time

t−1 through t−p and φ`,j(i) is the coefficient corresponding to the lag j value of the `-th channel Yt−`(j).

Thus, the VAR(p) model can be written as

Yt = ΦYt−1 + εt.

The least squares estimator for the unknown matrix Φ based on the observed time series {Yt, t = 1, . . . , n}is the minimizer

Φ = arg minΦ

d∑i=1

n∑t=1

[Yt(i)−Φ>(i)Yt−1]2.

The vector of estimators for the coefficients in the conditional mean of the i-th channel is then given by

Φ>

(i) =

(n∑t=1

Yt(i)Y>t−1

)(n∑t=1

Yt−1Y>t−1

)−1.

Let us define the residual vector to be Rt = Yt − ΦYt−1. If the null hypothesis of no change across

t = 1, . . . , n is true and that VAR model has a sufficiently large order providing a good approximation to

the covariance structure of Yt, then the expectation of Rt is (almost) zero and we can expect the partial

sum process to be “small”. Strong deviations of the partial sum process from zero, on the other hand,

indicate that the estimator of Φ does not provide a good fit to the data and could be evidence for a change

point. In fact, the first statistics considered for univariate AR time series were based on these partial sum

processes (see [Kulperger, 1985] and [Horvath, 1993]). However, such statistics can only detect certain

alternatives because they ignore other properties of the residuals such as the uncorrelatedness with past

observations. In particular, these statistics are meaningless if the time series are centered.

Consequently, [Davis et al., 1995] investigate the maximum likelihood statistic in the univariate AR

model, while [Huskova et al., 2007] consider the corresponding score type test and variations thereof.

The score type test is based not only on partial sums of the estimated residuals (in the non-centered

situation) but also on partial sums of the residuals multiplied with past observations which – under the

null hypothesis and correct specification – has also expectation zero, so that the partial sums should be

small. If this is not the case, the global estimator Φ does not fit the full data set well possibly due to

the existence of one or more change points and statistical inference based on this estimator is rendered

meaningless. All of the above statistics are particularly designed for the at-most-one-change situation.

3

Page 4: Detection of Changes in Multivariate Time Series With ...ckirch/seite/publications/media/pp_kmo_1… · Detection of Changes in Multivariate Time Series With Application to EEG data

While they do have some power against multiple changes, a fact binary segmentation procedures make

use of, their power is best in the at-most-one-change situation.

Tests particularly designed for the multiple change situation are only recently being developed but most

of them require at least an upper bound for the number of change points (see [Bai and Perron, 1998] and

[Marusiakova, 2009]. Exceptions are MOSUM based procedures discussed in [Huskova and Slaby, 2001]

and [Kirch and Muhsal, 2011]. On the other hand, there are established procedure available not for test-

ing but for estimation in a multiple change scenario that are based on the minimization of an appropriate

information criteria such as AutoParm developed by [Davis et al., 2006] which minimizes the minimum

description length in the context of autoregressive time series. However, these methods do not provide

measures of uncertainty (e.g., confidence levels) on the change point procedures. Similarly, tests particu-

larly designed for multivariate time series are only currently becoming of interest, some recent examples

are [Aue et al., 2009] and [Preußs et al., 2013].

In this paper, we extend the class of score type change point statistics considered by [Huskova et al., 2007]

to the vector autoregressive case and the epidemic change alternative. Furthermore, we no longer require

that the observed time series actually follows this model but take misspecification into account leading

to a detection procedure that uses the model background merely for feature extraction, something that

is implicitly done by many practitioners.

The remainder of the paper is organized as follows. In Section 2, we develop our statistical procedure

for testing and identification of change points in the at most one change (AMOC) and in the epidemic

settings. We derive the asymptotic distributions of our test statistics and show that the corresponding tests

have asymptotic power of 1. The proposed test statistics require the estimation of the inverse of the (long-

run) auto-covariance matrix a task particularly difficult in higher-dimensional settings. Consequently, in

Section 3, we describe methods of robustifying the proposed test statistics and investigate their finite

sample properties via extensive numerical experiments. We demonstrate the potential broad impact of

our procedure by analyzing electroencephalogram data set in Section 4.

2 Proposed Methods for Change Points Analysis

2.1 Test Statistics and Null Asymptotics

2.1.1 Background and Notation

Recall that the i-th channel in the VAR model is Yt(i) = Φ>(i)Yt−1+εt(i), where Φ(i) are the coefficients

for past data at all channels Yt−1. The number of coefficients for the i-th component is dp which is

typically large and consequently performing a large number of tests of hypotheses could lead to severe size

distortions or reductions in power (see [Aston and Kirch, 2013]) both of which can possibly be avoided

in situations where changes in the time series can be captured by a subset of the dp parameters. Of

course, there is always the danger of excluding a parameter that is sensitive to changes in the process.

Since our focus is on change point detection, we need to delicately balance between two competing aims:

(a.) ensuring a good approximation to the covariance structure which may require a VAR model with a

sufficiently high order and (b.) controlling the size of the test and ensuring sufficient power which requires

some careful reduction of the parameter space. Thus, our goal here is to provide a flexible procedure

that allows user-input by removing the VAR parameters that are determined, from prior knowledge,

4

Page 5: Detection of Changes in Multivariate Time Series With ...ckirch/seite/publications/media/pp_kmo_1… · Detection of Changes in Multivariate Time Series With Application to EEG data

to be not meaningful in change point detection. Moreover, since the d × dp matrix of parameters Φ

contains information on all lags of the covariance function cov(Yt+h,Yt), h = 0,±1, . . ., there will be

some redundant information shared by the elements of Φ. Hence, it is sensible to construct a test statistic

based only on a subset of Φ that captures the covariance features in the data well enough for change

detection purposes.

To develop our testing procedure, we first change the notation of the entries of Φ as follows

Φ(i) = [φ(i, 1), . . . , φ(i, p), . . . , φ(i, d(p− 1) + 1), . . . , φ(i, dp)]>

to simplify the description of the algorithm. Then, we define the indicator set I(i) = {r : φ(i, r) = 0 } to

be the coefficients, corresponding to the past values at all channels that are associated with the current

value for the i-th channel, that we set 0 for the change detection procedure and denote |I(i)| to be the

cardinality of this set. Thus, there remain dp − |I(i)| coefficients that are unknown and need to be

estimated. To obtain the specific past values of Yt−1 which correspond to the non-zero coefficients φ(i, r),

we introduce the following operator PI(·): for any dp-dimensional vector A = [A1, A2, . . . , Adp]>, let

PI(A) = [Ar, r 6∈ I]>

denote the vector, that only contains the elements not set 0 in the change detection procedure. Introducing

the notation

a(i) = PI(i) (Φ(i)) and Xt−1(i) = PI(i) (Ut−1)

with dim(Xt−1(i)) = dp− |I(i)| =dim(a(i)), the AR model (used in the change detection procedure) for

the i-th channel becomes Yt(i) = a(i)>Xt−1(i) + et(i). To emphasize the fact that we do not expect this

to be the true model, we rename the residuals {et} and allow them to be dependent. Then, the least

squares estimator for a(i) is given by

a>n (i) =

(n∑t=1

Yt(i)Xt−1(i)>)(

n∑t=1

Xt−1(i)Xt−1(i)>)−1

.

Example 2.1. One possible subset of coefficients that can be used to detect change points in multivariate

time series are the coefficients that relate Yt(i) with its own past observations Yt−1(i). This is appealing

because of interpretability as this is precisely what one would do if only the univariate information is

available for model building. It may appear at first glance that the auto-regressive parameters ignore

cross-dependence between channels of the multivariate time series. However, we demonstrate below that

this is not entirely true. For channel 1, the indicator set is I(1) = {(p+1), . . . , pd} and the set of non-zero

coefficients and corresponding past values are, respectively,

a(1) = [φ(1, 1), . . . , φ(1, p)]> and Xt−1(1) = [Yt−1(1), . . . , Yt−p(1)]>

where |I(1)| = dp − p = (d − 1)p; dim(Xt−1(1)) = p and dim(a(1)) = p. The AR model for channel 1

and for any channel i = 2, . . . , d are, respectively,

Yt(1) = φ1,1(1)Yt−1(1) + . . .+ φ1,p(1)Yt−p(1) + et(1)

= φ(1, 1)Yt−1(1) + . . .+ φ(1, p)Yt−p(1) + et(1),

Yt(i) = φ1,1(i)Yt−1(i) + . . .+ φ1,p(i)Yt−p(i) + et(i)

= φ(i, (i− 1)p+ 1)Yt−1(i) + . . .+ φ(i, ip)Yt−p(i) + et(i).

5

Page 6: Detection of Changes in Multivariate Time Series With ...ckirch/seite/publications/media/pp_kmo_1… · Detection of Changes in Multivariate Time Series With Application to EEG data

Note that the cross-dependence structure across channels is actually not being ignored in this change

point detection scheme because the within-channel (auto-) regression parameters capture information

concerning the cross-dependence structure and any left-over information will be absorbed by the residuals.

2.1.2 The At-Most-One-Change and Epidemic Change Point Settings

We shall develop a testing procedure for two types of change point situations: (1.) at-most-one-change

situation (AMOC) and (2.) epidemic change model which consists of two change points where the process

reverts back to the original regime after the second change point. Denote Y(1)t (i) to be the process at

regime Π1 and Y(2)t (i) at Π2. The AMOC model with a change point at k is given by

Yt(i) =

Y(1)t (i) = a1(i)

>X(1)t−1(i) + e

(1)t (i), 1 6 t 6 k,

Y(2)t (i) = a2(i)

>X(2)t−1(i) + e

(2)t (i), k < t 6 n.

(2.1)

The epidemic change model is given by

Yt(i) =

Y(1)t (i) = a1(i)

>X(1)t−1(i) + e

(1)t (i), 1 6 t 6 k1, k2 < t 6 n

Y(2)t (i) = a2(i)

>X(2)t−1(i) + e

(2)t (i), k1 < t 6 k2,

(2.2)

where the two change points are given by k1 and k2.

2.1.3 Assumptions on the Processes

Before we introduce the corresponding test statistics and their null asymptotics we first need to set

the stage by posting some assumptions on the underlying processes. In the correctly specified model

the innovations are i.i.d., but here we allow for misspecification which introduces some dependency to

the residuals. For example, if the true underlying process follows a full VAR-model but we only use

the restricted model of Example 2.1, the errors are no longer independent but contain the remaining

structural information of the time series. The assumptions below allow for this but also for other kinds

of misspecification.

Assumption A. 1. Let {Y(1)t = (Y

(1)t (1), . . . , Y

(1)t (d))> : t > 1}, be an R-valued strictly stationary

sequence of non-degenerate random vectors with

E Y(1)1 = 0, E ‖Y(1)

1 ‖4+ν <∞ for some ν > 0,

satisfying a strong mixing condition with mixing rate

α1(n) = O(n−β

)for some β > max

(3,

4 + ν

ν

).

We merely make this classic mixing assumption for simplicity of presentation because it ensures

invariance principles for functionals of the time series. However, our results carry over to different more

recent weak dependency concepts as long as the necessary invariance principles are still available. In the

correctly specified case, this assumption is not necessary (e.g. one can allow for discrete errors) as other

tools to derive invariance principles are available (for details we refer to [Huskova et al., 2007]).

6

Page 7: Detection of Changes in Multivariate Time Series With ...ckirch/seite/publications/media/pp_kmo_1… · Detection of Changes in Multivariate Time Series With Application to EEG data

A large class of these weak dependent processes such as ARMA-models or linear processes (in the

strict sense) can well be approximated by a large enough linear autoregressive model as given in (2.1).

Define

a1(i) = C−11 (i)c1(i), e(1)t (i) = Y

(1)t (i)− a>1 (i)X(1)

t−1(i),

C1(i) = E(X(1)t−1(i)(X

(1)t−1(i))

>), c1(i) = E

(X(1)t−1(i)Y

(1)t (i)

). (2.3)

By Assumption 1 and Davydov’s covariance inequality (see [Kuelbs and Philipp, 1980], Lemma 2.1)

the autocovariance function γ(h) of Y(1)t converges to zero as h → ∞, hence C1(i) is invertible (see

[Brockwell and Davis, 1991], Proposition 5.1.1). The parameter vector a1(i) is best approximating in the

following sense:

a1(i) = arg minb1(i)

E

((Y

(1)t (i)− b1(i)

>X(1)t−1(i)

)2),

so that the parameters in (2.3) are equal to the true causal parameters in the correctly specified case.

Under alternatives, we need to make the following assumptions:

Assumption A. 2. Let {Y(2)t = (Y

(2)t (1), . . . , Y

(2)t (d))> : t > 1}, be an R-valued ergodic sequence of

random vectors with existing second moments and

1

n− k

n∑j=k+1

E(X(2)t−1(i)(X

(2)t−1(i))

>)→ C2(i),1

n− k

n∑j=k+1

E(X(2)t−1(i)Y

(2)t (i))→ c2(i)

in case of the AMOC alternative respectively

1

k2 − k1

k2∑j=k1+1

E(X(2)t−1(i)(X

(2)t−1(i))

>)→ C2(i),1

k2 − k1

k2∑j=k1+1

E(X(2)t−1(i)Y

(2)t (i))→ c2(i)

in case of an epidemic change alternative.

This assumption obviously holds for stationary sequences, however it also allows for starting values

from the stationary distribution of Y(1). Similarly, in case of the epidemic change alternative, we can

allow the process after the second change point to have starting values from the distribution of Y(2).

2.1.4 Test Statistics for the At-Most-One-Change Situation

For the AMOC situation, we propose the following class of test statistics:

M (1)n = max

16k6n

w2(k/n)

nZ>k HZk, (2.4)

M (2)n =

∑16k6n

w2(k/n)

n2Z>k HZk (2.5)

Z>k =(S>k (1), . . . ,S>k (d)

), Sk(i) =

k∑t=1

ξt(i) =

k∑t=1

Xt−1et(i),

et(i) = Yt(i)− a>n (i)Xt−1(i),

an(i) = C−1n (i)1

n

n∑i=1

Xt−1(i)Yt(i), Cn(i) =1

n

n∑t=1

Xt−1(i) (Xt−1(i))> . (2.6)

7

Page 8: Detection of Changes in Multivariate Time Series With ...ckirch/seite/publications/media/pp_kmo_1… · Detection of Changes in Multivariate Time Series With Application to EEG data

The matrix H is pd× pd symmetric and positive semi-definite and fulfills HP−→H for some symmetric,

positive semi-definite matrix H under H0. Possible choices of H and w(·) are discussed in Section 2.1.6

below.

The above test statistics also have some power against multiple change point situations including

epidemic changes which is the basic idea behind binary segmentation algorithms. The point, where the

maximum is attained, is a consistent estimator for the change point in rescaled time (see Section 2.2.2

for details). In the EEG data example, we apply this estimator for each of the independent trials and

plot the corresponding density estimates, which appear to be bimodal (see Figure 4.2). Therefore, we

suspect that the EEG process returned to the “waiting” state after the task (moving the joystick) has

been completed, which is why the following test statistics for epidemic changes are of particular interest

for that data example.

2.1.5 Test Statistics for the Epidemic Situation

The following test statistics are designed to detect epidemic changes. See [Csorgo and Horvath, 1997]

(Section 2.8.4) for corresponding statistics for mean changes.

M (3)n = max

bτ1nc6k1<k26n−bτ2nc

1

n(Zk2 − Zk1)>H(Zk2 − Zk1), (2.7)

M (4)n =

1

n3

∑bτ1nc6k1<k26n−bτ2nc

(Zk2 − Zk1)>H(Zk2 − Zk1), (2.8)

where 1 6 τ1 6 1− τ2 6 1 and the rest of the notation is as in(2.6). Here, we only allow for very simple

weight functions

w(t1, t2) = 1{τ16t1,t261−τ2}

because the asymptotics in the epidemic setting with more general weight functions can become more

involved. Similarly, to the at most one change situation, we can estimate both change points (in rescaled

time) by the points where the maximum is attained (for details we refer to Section 2.2.2).

2.1.6 Some Remarks on the Tuning Parameters H and w

To understand the influence of H better, we define

ξt = (ξ>t (1), . . . , ξ>t (d))> =(

(X(1)t−1(1))>e

(1)t (1), . . . , (X(1)

t−1(d))>e(1)t (d)

)>(2.9)

and note that the asymptotic covariance of the partial sum process Zk is given by the covariance of ξ in

the correctly specified model and by its long-run covariance in the misspecified case (confer Theorem 2.1

for details).

Now, the choice of the weight matrix H allows us to look for changes in particular directions more

intensely. The following choice, e.g., leads to the univariate statistics, introduced by [Huskova et al., 2007],

that looks for changes only in the first channel of the observed multivariate time series:

H =

(cov−1(ξ0(1)) 0

0 0

)(2.10)

8

Page 9: Detection of Changes in Multivariate Time Series With ...ckirch/seite/publications/media/pp_kmo_1… · Detection of Changes in Multivariate Time Series With Application to EEG data

The standard choice for H in the correctly specified case is given by

H = cov−1 (ξ0) . (2.11)

However, if the channels of the observations are correlated the estimation of this matrix let alone its

inverse is very difficult in higher dimensional settings (even for moderate dimensions if there are not that

many time points available). Things become even worse if we are in the misspecified situation where

the inverse of the long-run variance needs to be estimated. Furthermore, in the change point situation

possible alternatives should be taken into account in the estimation of H in order to improve the power

of the corresponding test, making things even more complicated. So, some practical solutions to this

problem will be discussed in detail in Section 3.

The weight function w(·) essentially determines where we look closest for a change since the statistic

has a higher power for changes close to k if w(k/n)/cw is larger where cw is the critical value associated

with w(·). While the matrix H can be used to incorporate a priori information about the direction of

the expected changes, we can use the weight function w(·) to incorporate a priori information about the

location of the change points. In the EEG data example we do not expect a change before the signal was

sent, so we will choose a weight function that only looks for changes in the second half of the data set.

To derive the asymptotics for the AMOC statistic, we need to formulate the following assumptions:

Assumption A. 3. The weight function w : [0, 1] → [0,∞) is a non-negative function, w 6≡ 0. Further-

more, it has only a finite number of discontinuities a1, . . . , aK , where it is either left or right continuous

with existing limits from the other side. Additionally, the following regularity conditions hold

limt→0

tαw(t) <∞, limt→1

(1− t)αw(t) <∞ for some 0 6 α < 1/2,

supη6t61−η

w(t) <∞ for all 0 < η 61

2.

Typical symmetric weight functions are given by

w1(t) = 1{ε<t<(1−ε)}(t(1− t))−1/2,

w2(t) = (t(1− t))−β for some 0 6 β < 1/2.

2.1.7 Asymptotic Results under the Null

We are now ready to state the asymptotic distribution of the proposed test statistics under the null

hypothesis.

Theorem 2.1. Under the null hypothesis let the process {Yt = Y(1)t } fulfill A.1.Let the weight function

fulfill A.3 and HP−→ H (under H0), then

(a) M (1)n

D−→ sup06t61

w2(t)

pd∑j=1

B2j (t),

(b) M (2)n

D−→∫ 1

0w2(t)

pd∑j=1

B2j (t) dt,

9

Page 10: Detection of Changes in Multivariate Time Series With ...ckirch/seite/publications/media/pp_kmo_1… · Detection of Changes in Multivariate Time Series With Application to EEG data

where {B(t) = (B1(t), . . . , Bpd(t))> : t ∈ [0, 1]} denotes a pd-dimensional Brownian bridge with covariance

matrix H12 ΣH

12 . Σ is the long-run autocovariance matrix of {ξt} as in (2.9), i.e. Σ =

∑k∈Z E ξ0ξ

>k .

Theorem 2.2. Under the null hypothesis let the process {Yt = Y(1)t } fulfill A.1.Let H

P−→ H (under

H0), then

a) M (3)n

D−→ supτ16t1<t261−τ2

pd∑j=1

(Bj(t2)−Bj(t1))2 ,

b) M (4)n

D−→∫ ∫

τ16t1<t261−τ2

w2(t1, t2)

pd∑j=1

(Bj(t2)−Bj(t1))2dt1 dt2,

where {B(t) = (B1(t), . . . , Bpd(t))> : t ∈ [0, 1]} denotes a pd-dimensional Brownian bridge with covariance

matrix H12 ΣH

12 . Σ is the long-run autocovariance matrix of {ξt} as in (2.9), i.e. Σ =

∑k∈Z E ξ0ξ

>k .

2.2 Tests and Estimators under Alternatives

In this section, we will show that the above statistics detect a large class of alternatives asymptotically.

In fact, in the correctly specified case, all alternatives are detectable if a positive definite weight matrix

H is used. In this situation the corresponding estimators are consistent.

We make the following standard assumptions on the locations of the change point:

Assumption A. 4. For the at most one change situation we assume that k = bλnc, 0 < λ < 1, for the

epidemic change alternative we assume kj = bλjnc, 0 < λj < 1, j = 1, 2.

In order to classify detectable changes under misspecification we need to understand the behavior

of the estimator an as in (2.6) under the alternative. In this situation it can no longer be expected to

converge to the true or best approximating value before the change but it will be contaminated by the

alternative. In fact, Lemma 6.1 shows that an(i) converges to

a(i) = Q−1(i)q(i), i = 1, . . . , d,

where Q(i) = λC1(i) + (1− λ)C2(i), q(i) = λc1(i) + (1− λ)c2(i),

λ =

λ, at most one change,

1− (λ2 − λ1), epidemic change.(2.12)

The matrix Q(i) is the weighted average of the autocovariances of the time series, before and after the

change, with weights proportional to the duration of time the series remains in that state. Here, λ is the

duration time in state 1. We now make the additional assumption:

Assumption A. 5. Let Q(i), i = 1, . . . , d, be regular matrices.

This assumption guarantees that a(i), i = 1, . . . , d, above are well defined and the unique minimizer

of

(1− λ) E(Y(1)1 (i)− a>(i)X(1)

1 (i))2 + λE(Y(2)1 (i)− a>(i)X(2)

1 (i))2.

10

Page 11: Detection of Changes in Multivariate Time Series With ...ckirch/seite/publications/media/pp_kmo_1… · Detection of Changes in Multivariate Time Series With Application to EEG data

2.2.1 Asymptotic Power One under Alternatives

Theorem 2.3. Let Assumptions A.1, A.2, A.3, as well as A.5 hold true. Furthermore, assume that there

exists ε > 0 such that

P (∆2H

> ε)→ 1, (2.13)

where ∆2H

= (c1 − diag(C1(1), . . . ,C1(d))a)> H−1

(c1 − diag(C1(1), . . . ,C1(d))a) .

a) Under the at most one change alternative, it holds

(i) M (1)n

P−→∞, (ii) M (2)n

P−→∞.

b) Under the epidemic change alternative, it holds

(i) M (3)n

P−→∞, (ii) M (4)n

P−→∞.

Remark 2.1. Condition (2.13) is the heart of the theorem explaining which changes are detectable.

If HP−→ HA for a positive definite matrix, then all changes are detectable for which C−11 (i)c1(i) 6=

C−12 (i)c2(i) for at least one of the channels i = 1, . . . , d. Hence, asymptotically all changes in the correctly

specified model are detectable. If HA has a block diagonal form with block lengths of p such that the i-th

block matrix is positive definite, then we will detect all changes that appear in channel i, i.e. for which

C−11 (i)c1(i) 6= C−12 (i)c2(i).

2.2.2 Estimators for the Change Point(s) under Alternatives

If the test statistics are consistent we also obtain consistent estimators as follows: Under the AMOC

model, we define the change point estimator as

kn := arg max1<k<n

w2(k/n)STk HSk (2.14)

and for the epidemic model we define(k1,n, k2,n

):= arg max

τ16k1<k26τ2(Sk2 − Sk1)>H(Sk2 − Sk1). (2.15)

Theorem 2.4. Let the assumptions of Theorem 2.3 be fulfilled in addition to H →HA, where HA can

be different from H.

(a) Under the AMOC change alternative assume additionally that sup06t61w(t) < ∞ and let the

function

h(t) = w(t)

t(1− λ), t 6 λ,

(1− t)λ, t > λ

has a unique supremum at λ, which holds e.g. for the weight functions w(t) = (t(1 − t))−β,

0 6 β < 1/2. Then

knn

P−→ λ.

11

Page 12: Detection of Changes in Multivariate Time Series With ...ckirch/seite/publications/media/pp_kmo_1… · Detection of Changes in Multivariate Time Series With Application to EEG data

(b) Under the epidemic alternative assume additionally that τ1 < λ1 < λ2 < τ2, then(k1,nn,k2,nn

)P−→ (λ1, λ2) .

Remark 2.2. If we want to use more general weight functions as given in Assumption A.3 for the at

most one change alternative then we need to make the additional assumption that the time series after

the change {Y (2)t (l), l = 1, . . . , d} is also strong mixing with a rate as given in Assumption A.1.

3 Robustifying the Test Statistics and Empirical Study

In order to use asymptotic critical values in the above test statistics we need to choose H (resp. H) in

such a way that the limit becomes pivotal. The standard solution is to choose H = Σ−1 as in (2.11)

where Σ is the long-run covariance matrix of {ξt} as in (2.9). For most applications the true underlying

long-run variance (under H0) will not be known so it needs to be estimated. However, this raises several

problems:

P.1 When designing the estimation procedure, we do not only need it to be consistent under the null

hypothesis, but also to behave reasonably under alternatives e.g. converging to some contaminated

limit matrix HA. In fact, the small sample power behavior will depend crucially on this choice.

P.2 Even in the univariate situation it is a much harder statistical problem to estimate the long-run

variance than the variance. Consequently, the estimation error for the long-run variance is much

bigger than for the variance (for fixed sample size).

P.3 Increasing the number of entries of H that need to be estimated leads to more imprecise estimates.

Typically, we need to estimate some statistical quantity such as the covariance or long-run covariance

matrix and then invert the estimated matrix to obtain H which leads to even less precise estimates and

additional numerical errors. As a consequence, the size can be very distorted (see Figure 3.1 below). A

block diagonal structure and many zeros in H can help overcome these estimation issues and stabilize the

empirical size.

3.1 Robustifying the Univariate Statistic

3.1.1 Stabilizing the Power

The univariate situation is not only included by d = 1 but also by choosing Xt−1(1) = (Yt−1(1), . . . , Yt−p(1))>

and letting

H =

(H(1) 0

0 0

).

In this section, we choose the second approach because it makes generalizations to the multivariate

situation in the next section more transparent.

Ideally, we choose H(1) in such a way that (a.) the limit becomes pivotal under the null hypothesis

and (b.) all parameter changes in the correctly specified model respectively all changes leading to different

12

Page 13: Detection of Changes in Multivariate Time Series With ...ckirch/seite/publications/media/pp_kmo_1… · Detection of Changes in Multivariate Time Series With Application to EEG data

best approximating parameters under misspecification should be asymptotically detectable. We achieve

both goals by choosing the inverse of the covariance matrix of {ξt(1)} in the correctly specified and of the

long-run covariance matrix of {ξt(1)} in the misspecified case. However, {ξt(1)} is not observable, hence

we need to use estimated versions which because of P.1 should be chosen in such a way that they take

alternatives into account. Under the at most one change alternative typically the following approach is

taken (see, e.g., [Huskova et al., 2007]): Use a preliminary estimator H for H which does not take the

alternative into account. Estimate the change point k(1) as

k(1) = arg maxw2(k/n)

nZ>k HZk,

then estimate the residuals before and after the change by

ξ(H1)t (1) = Xt−1(1)et(1), t = 1, . . . , n

where et(1) =

Yt(1)− Xt−1(1)>ak(1)

(1), t 6 k(1),

Yt(1)− Xt−1(1)>a◦k(1)

(1), t > k(1).

ak(1)

(1) is the least squares estimator for a(1)(1) based on Y1(1), . . . , Yk(1)

(1) while a◦k(1)

(1) is the least

squares estimator for a(2)(1) based on Yk(1)+1

(1), . . . , Yn(1). For the epidemic change alternative an

analogous version can be used. This approach works quite well, if the alternative is correctly specified

and the residuals before and after the change have similar statistical properties. Otherwise, e.g. in the

presence of multiple changes, a systematic estimation error is introduced. While this systematic error is

typically asymptotically negligible as long as the limit is sufficiently nice (see Section 2.2), it usually leads

to some power loss in small samples.

3.1.2 Stabilizing the Power for the EEG Data

By design, we do not expect pre-stimulus changes in the EEG time series (i.e., during the first half of

each trial). Following stimulus presentation, there may be gradual changes or multiple changes. First,

there could be brief changes in the EEG due to the visual stimulus followed by changes that accompany

brain processing of the visual information. Consequently, the approach discussed in the previous section

does not seem ideal here. However, because the data set is designed in such a way that the first half is

stationary, we can and will only use the data set up to time point 250 in the estimation thus automatically

taking possible change points after the visual signal into account. Precisely, we will use

ξ(250)t (1) = Xt−1(1)et(1), t = 1, . . . , 250, where et(1) = Yt(1)− Xt−1(1)>a250(1),

and a250(1) is the least squares estimator for a(1)(1) based only on Y1(1), . . . , Y250(1).

3.1.3 Stabilizing the Size with Respect to Possible Misspecification

Based on the estimated residuals from Sections 3.1.1 or 3.1.2 the inverse of the long-run covariance matrix

can be estimated e.g., using flat-top kernel estimators with an automatic bandwidth selection criteria (as

proposed in [Politis, 2003]). However, even in the one-dimensional case where only the long-run variance

13

Page 14: Detection of Changes in Multivariate Time Series With ...ckirch/seite/publications/media/pp_kmo_1… · Detection of Changes in Multivariate Time Series With Application to EEG data

needs to be estimated the estimation error can become quite large. Therefore, we propose to trade this

possibly large estimation error for a hopefully smaller model error. To this end, we propose to increase

the length p of the fitted AR-model which can approximate a large class of time series quite well so that

the corresponding errors behave more like i.i.d. time series even in the misspecified case. This allows us

to use an estimator for the covariance matrix rather than the long-run covariance matrix.

A possible drawback is that this approach now increases the unknown number of parameters of the

weight matrix H associated with problem P.3 above. Additionally, we now need to look for changes in

a larger number of parameters which in combination with the choice H(1) = cov ξ1(1) leads to a power

loss if this change is actually already quite visible in the smaller model. This is why, we propose to only

look for particular changes by choosing

H(1) =

(cov

[e(1)t (1) (Y

(1)t−1(1), . . . , Y

(1)t−p′(1))>

]0

0 0

), (3.1)

for some p′ < p. This choice also leads to a pivotal limit.

3.1.4 Investigating the Performance of the Robustified Univariate Procedures

To study the effect of these robustifications, we generate data from the following AR(6) time series (with

i.i.d. standard normally distributed errors and a change at t = 330):

• Case AR1. a1(1) = (−0.1, 0.1, 0, 0,−0.1,−0.1)>, a2(1) = (−0.4, 0.1, 0, 0,−0.1,−0.1)>, i.e. the

change occurs only in the first parameter. Moreover, the last four channels of a1(1) are small.

Thus, by fitting an AR(2) model to the data, our procedure only suffers from mild model misspeci-

fication.

• Case AR2. a1(1) = (−0.1, 0.1, 0, 0.2,−0.3,−0.2)>, a2(1) = (−0.2, 0.3, 0, 0.2,−0.3,−0.2)>, i.e. the

change occurs only in the first two parameters but there is stronger misspecification when using the

AR(2) model.

• Case AR3. a1(1) = (−0.1, 0.1, 0, 0,−0.3,−0.2)>, a2(1) = (−0.1, 0.2, 0, 0.4,−0.1,−0.2)>, i.e. there is

only a small change in the second parameter in addition to changes in the fourth and fifth parameter

again under a stronger misspecification when using the AR(2) model.

In all cases, we used the following weight function adapted to the EEG data example w(t) = I[ 250512, 490512 ](t)(1−

t)−0.25, t ∈ [0, 1].

Figure 3.1 gives the results for several choices of p and p′ and the maximum type statistic M(1)512 as in

(2.4) as well as the sum type statistic M(2)512 as in (2.5) with H as described in Sections 3.1.2 and 3.1.3.

The first row shows the empirical size on the y-axis for the nominal one on the x-axis, the second row

shows the size-corrected power. For p = p′ = 2 we additionally use both the covariance estimator based on

the first 250 observations (C250) as well as the flat-top kernel estimator in [Politis, 2003] with automatic

bandwidth selection choice (S250). Clearly, the size behavior is best in all three cases for p = 6, p′ = 2

as we expected and this is the only choice with an acceptable size behavior. For p = 2, p′ = 2 the size

behavior is not acceptable due to a too large an estimation error (when estimating the long-run covariance

matrix) and too large a model error (when estimating the covariance matrix) and is worse the stronger

14

Page 15: Detection of Changes in Multivariate Time Series With ...ckirch/seite/publications/media/pp_kmo_1… · Detection of Changes in Multivariate Time Series With Application to EEG data

0.00 0.05 0.10 0.15 0.20

0.00

0.10

0.20

0.30

0.00 0.05 0.10 0.15 0.20

0.0

0.2

0.4

0.6

0.8

1.0

S250, p=2, p'=2C250, p=2, p'=2C250, p=6, p'=6C250, p=6, p'=2

M1M2

0.00 0.05 0.10 0.15 0.20

0.00

0.10

0.20

0.30

0.00 0.05 0.10 0.15 0.20

0.0

0.2

0.4

0.6

0.8

1.0

S250, p=2, p'=2C250, p=2, p'=2C250, p=6, p'=6C250, p=6, p'=2

M1M2

0.00 0.05 0.10 0.15 0.20

0.00

0.10

0.20

0.30

0.00 0.05 0.10 0.15 0.20

0.0

0.2

0.4

0.6

0.8

1.0

S250, p=2, p'=2C250, p=2, p'=2C250, p=6, p'=6C250, p=6, p'=2

M1M2

Figure 3.1: Empirical size and size-corrected power for AR1 – AR3 (top to bottom). C250 uses the covariance

estimate; S250 uses the long run covariance estimate based on the first 250 observations.

the misspecification. For p = 6, p′ = 6 there is no model error but nevertheless the estimation error due

to P.3 makes the tests far too liberal (although not as badly so as for the choice p = p′ = 2). Using the

sum-type statistic is in all cases more robust and gives a better size while not losing actual power. In

the first two cases, where the main change occurs in the first two parameters, the power for the choice

p = 6, p′ = 2 is best, even better than for p = 2, p′ = 2, while also being good for this second choice. As

expected we do lose some power in this case for p = 6, p′ = 6. However, if the change mainly occurs in

the last four parameters as in the third case above, the power for p = 6, p′ = 6 is better than for the other

cases while they still have some power against this alternative. The bootstrap procedure of Section 3.2.1

will further improve the size.

3.2 Robustifying the Multivariate Statistics

In general, if a change is present in (almost) all channels, one obtains higher power when the information

is pooled and the multivariate statistic is used. The statistic at hand is a quadratic form so that the

15

Page 16: Detection of Changes in Multivariate Time Series With ...ckirch/seite/publications/media/pp_kmo_1… · Detection of Changes in Multivariate Time Series With Application to EEG data

changes in each channel add up quadratically, while the variance of the noise (if reasonably uncorrelated)

adds up only linearly, leading to a better signal-to-noise ratio of the multivariate statistic. On the other

hand, if there is no change in many channels then we merely add noise and thus decreasing the power.

If additional information about the direction of the change is available appropriate projections (which

are included in the above theory by an appropriate choice of H) can further help increase the power. A

detailed analysis of this effect in the simpler mean change model can be found in [Aston and Kirch, 2013].

3.2.1 Size Correction: A Bootstrap Approach

In the multivariate procedure, the estimation problem becomes even worse because the number of entries

of H that need to be estimated increases polynomially with the number of dimensions. Since we already

have massive estimation problems in the univariate case, estimating the inverse of the covariance or even

long-run covariance matrix of {ξt} is statistically infeasible. To overcome this problem, we first choose to

relate Yt(i) only to its own past as in Example 2.1 and secondly to choose

H = diag(H(1),H(2), . . . ,H(d)) (3.2)

with H(j) analogously to (3.1). In case the channels were independent this would lead to a pivotal limit.

However, we do not believe that the channels are independent, hence the limit distribution depends on that

underlying covariance structure which we cannot estimate in a stable way. As a solution, we propose to use

the following regression bootstrap (which is related to the one investigated in [Horvath, L. et al., 1999]).

We also implemented a version of the pair bootstrap that was discussed in that paper, but the regression

bootstrap was always superior in simulations. The bootstrap works as follows:

(1) Calculate the residuals e1(j), . . . , e250(j), j = 1, . . . , 250:

et(j) = Yt−1(j)− XTt−1(j)a250(j)−1

250

250∑i=1

(Yi(j)− XTi−1(j)a250(j)

), 1 6 t 6 250,

where a250(j) is the least squares estimator based on Y1(j), . . . , Y250(j).

(2) Draw n i.i.d. random variables U1, . . . , Un such that P (U1 = i) = 1/250, i = 1, . . . , 250.

(3) Let for k = 1, . . . , n, j = 1, . . . , d

e∗k(j) := eUk(j)

and ξ∗t :=(XTt (1)e∗t (1), . . . ,XTt (d)e∗t (d)

)T, t = 1, . . . , n.

(4) Calculate the multivariate statistics from the previous sections, but with Zk replaced by Z∗k and H

as in (3.2) by H∗ where

Z∗k =k∑i=1

ξ∗i − CkC−1n

n∑i=1

ξ∗i ,

Ck = diag(Ck(1), . . . , Ck(d)), Ck(i) =1

k

k∑t=1

Xt−1(i)XTt−1(i)

16

Page 17: Detection of Changes in Multivariate Time Series With ...ckirch/seite/publications/media/pp_kmo_1… · Detection of Changes in Multivariate Time Series With Application to EEG data

and H∗n is a block diagonal matrix with j = 1, . . . , d blocks (H∗n(j))−1, where

H∗n(j) =

1250

∑250l=1

(e∗l (j)−

1250

∑250s=1 e

∗s(j)

)21

250

∑250i=1 Xi−1(j)XTi−1(j) 0

0 0

,

Xi−1(j) = (Xi−1(j), . . . , Xi−p′(j))T .

(5) Repeat steps (2)-(4) M times (M = 2000).

(6) The critical value c∗(α) is obtained as the upper α-quantile of the M statistics.

(7) Reject H0 if the corresponding statistic based on the original sample exceeds the critical value c∗(α).

3.2.2 Investigating the Performance of the Robustified Multivariate Procedure

To investigate the influence of dependency between panels on the asymptotic procedure and the per-

formance of the proposed bootstrap method, we consider channelwise AR-models, where there is some

correlation between the errors among channels. In each channel the following models are used: a1 =

(0.5,−0.2, 0.1, 0, 0, 0.2)> before the change and a2 = (0.5,−0.3, 0.1, 0, 0, 0.2)> after the change. The

multivariate errors are normal with mean zero and variance 1 and given as follows:

E.1 The errors are uncorrelated.

E.2 There is a correlation of 0.2 between any two channels.

E.3 The correlation is 0.6 between any two channels.

E.4 There is a correlation of 0.1 between any two channels except for two pairs of channels where the

correlation is 0.5.

E.5 We additionally have three pairs with a correlation of 0.3.

The empirical results can be found in Figure 3.2. In all cases the bootstrap performs very well in terms of

size and only slightly inferior in terms of power showing that it is clearly robust against deviations from

independence between channels. The size of the asymptotic tests on the other hand are not so good and

in particularly strongly distorted for case 3 where there is a strong correlation between all panels.

3.3 Sensitivity Study under Alternatives

In this section, we study the sensitivity of the procedures with respect to the location of the change, the

epidemic duration length, the size of changes and the number of channels affected by change points. With

a view to the EEG data we choose AR(8) time series of length 512 with various epidemic changes, i.e.

the univariate time series are given by

Yt(1) = a1(1)>(Yt−1(1), . . . , Yt−1(8)) + 1{t16t<t2}d(1)>(Yt−1(1), . . . , Yt−1(8)) + et, (3.3)

where {et} are i.i.d. standard normally distributed and model parameters a1 = (0.5, 0, 0.1, 0, 0, 0.2, 0.1,−0.2)>,

a1 +d(1) = (0.5, a22, 0.1, 0, 0, 0.2, 0.1,−0.2)>. We apply the epidemic change statistics M(3)512 and M

(4)512 as

17

Page 18: Detection of Changes in Multivariate Time Series With ...ckirch/seite/publications/media/pp_kmo_1… · Detection of Changes in Multivariate Time Series With Application to EEG data

0.00 0.05 0.10 0.15 0.20

0.00

0.10

0.20

0.30

0.00 0.05 0.10 0.15 0.20

0.00

0.10

0.20

0.30

0.00 0.05 0.10 0.15 0.20

0.00

0.10

0.20

0.30

0.00 0.05 0.10 0.15 0.20

0.00

0.10

0.20

0.30

0.00 0.05 0.10 0.15 0.20

0.00

0.10

0.20

0.30

0.00 0.05 0.10 0.15 0.20

0.0

0.2

0.4

0.6

0.8

1.0

0.00 0.05 0.10 0.15 0.20

0.0

0.2

0.4

0.6

0.8

1.0

0.00 0.05 0.10 0.15 0.20

0.0

0.2

0.4

0.6

0.8

1.0

0.00 0.05 0.10 0.15 0.20

0.0

0.2

0.4

0.6

0.8

1.0

0.00 0.05 0.10 0.15 0.20

0.0

0.2

0.4

0.6

0.8

1.0

M1 asymp.M2 asymp.M1 boot.M2 boot.

Figure 3.2: Empirical size (top) and size-corrected power (bottom) with dependent errors according to E.1 – E.5

(left to right)

in (2.7) and (2.8) with weight function 1{250/5126t6490/512} (as will be used in the EEG data due to the

prior knowledge that there should not be a change before point 250 or after 490). The weight matrix H

will be chosen according to (3.2) with p = 6, p′ = 2 (allowing for some misspecification). We do not apply

the bootstrap due to computational time but report the size corrected power to take size variations into

account.

The empirical size corrected power to the 5% level for a change in a single parameter based on

2000 repetitions is reported in Table 3.1 Obviously, the larger the change and the longer the epidemic

duration, the better the power, where changes to a negative parameter are apparently more difficult to

detect. Nevertheless, the power is quite good even for small epidemic durations (30 out of 512 observations

are in the epidemic state) and can further be increased by using the full multivariate information.

To demonstrate the power behavior under the multivariate setting, we simulate i.i.d. channels as in

(3.3) with a22 = −0.4. We allow the change locations to slightly vary across channels by independently

drawing the change points for each channel. To this end we choose uniform distributions centered around

300 and 350 respectively with several ranges±0,±5 and±10. Table 3.2 reports the empirical size corrected

power for the asymptotic 5% level. As expected, Table 3.2 shows clearly that the power increases if the

multivariate statistics are used and changes are present in several channels. Furthermore, it shows that

the power decreases if only noise is added. Interestingly, the power does not seem to be affected by a slight

variation of change locations across channels, a situation which one would expect in many applications.

4 Data Analysis

We analyze data consisting of electroencehpalograms (EEG) recorded from a single participant in a

visual-motor task experiment. Our neuroscientist collaborator is broadly interested in studying brain

dynamics during sensory information and execution of the motor task. In this study, we tackle two

primary statistical goals. The first is to identify the change points in brain signals within a trial. The

second is to study how the change point structure varies across many trials across the experiment. The

neuroscience community is currently highly interested in characterizing variations in brain responses. To

the best of our knowledge, our procedure is the first to systematically investigate between-trial variation

18

Page 19: Detection of Changes in Multivariate Time Series With ...ckirch/seite/publications/media/pp_kmo_1… · Detection of Changes in Multivariate Time Series With Application to EEG data

5% a22 = 0.6 a22 = 0.4 a22 = 0.2 a22 = −0.2 a22 = −0.4 a22 = −0.6

t1 = 300, t2 = 330 0.973 0.7350 0.152 0.055 0.1465 0.3875

0.970 0.7285 0.148 0.067 0.1590 0.4325

t1 = 300, t2 = 350 0.9995 0.9245 0.257 0.0725 0.292 0.7245

0.9990 0.9235 0.267 0.0910 0.338 0.7655

t1 = 300, t2 = 400 1 0.9980 0.515 0.1635 0.7130 0.9875

1 0.9985 0.557 0.2070 0.7905 0.9935

Table 3.1: Size-corrected power for 5% level and univariate versions of the statistics M(3)512 (upper value)

and M(4)512 (lower value).

in change points.

4.1 Data Description

The EEG data are electric potentials recorded from the scalp and thus reflect indirect measures of brain

electrical activity. In this experiment, the participant was instructed to move the hand-held joystick either

towards the left of center or towards the right of center. There were N1 = 118 leftward movement trials

and N2 = 134 rightward movement trials. The order of presentation of the left and right conditions was

random.

The EEG was recorded at the rate of 512 Hz and band-pass filtered at (0.02, 100) Hz. Each of the

N1+N2 = 252 trials consists of one second (total of n = 512 time points) whose midpoint is approximately

the point in time when the stimulus was presented (at t = 250). Thus, each trial consists of half a second

of EEG recording before stimulus presentation and another half a second post stimulus presentation.

From the montage of 64 scalp electrodes, our neuroscientist collaborator selected a sub-set of 12 surface

leads at: FC3, FC5, C3, P3, and O1 over the left hemisphere; FC4, FC6, C4, P4 and O2 over the right

hemisphere; and Cz and Oz over the mid-line. The frontal (FC) leads were presumably placed over the

prefrontal cortex, regions previously shown to have involvement in premotor processing. The central (C)

leads were placed over structures involved in motor performance, while the parietal (P) and occipital

5% 3/4 2/4 1/4 6/8 4/8 2/8 9/12 6/12 3/12

±0 0.49 0.31 0.15 0.89 0.74 0.22 0.88 0.61 0.27

0.59 0.41 0.20 0.94 0.83 0.32 0.94 0.78 0.42

±5 0.49 0.32 0.15 0.88 0.73 0.22 0.88 0.61 0.28

0.59 0.41 0.20 0.94 0.83 0.32 0.94 0.78 0.41

±10 0.48 0.33 0.16 0.87 0.71 0.23 0.86 0.61 0.28

0.59 0.40 0.20 0.93 0.82 0.31 0.93 0.75 0.39

Table 3.2: Size-corrected power for 5% level and statistics M(3)512 (upper value) and M

(4)512 (lower value)

when x out of y (x/y) channels did contain epidemic change points which were located independently for

each channel uniformly around 350 and 380 respectively according to ±z.

19

Page 20: Detection of Changes in Multivariate Time Series With ...ckirch/seite/publications/media/pp_kmo_1… · Detection of Changes in Multivariate Time Series With Application to EEG data

(O) leads were placed over structures involved in visual sensation and visual motor transformations (see

[Marconi et al., 2001]).

4.2 Data Analysis

To determine optimal order of the VAR for this EEG data, we applied both the Akaike information

criterion (AIC) and the Bayesian information criterion (BIC) for each trial using only the pre-stimulus

period because this period, consisting of the first 250 time points, is approximately stationary. Both AIC

and BIC produced optimal order d that ranged from 9 to 17 across the many trials. Using the lowest order

of d = 9 already produced residuals whose auto-correlation and partial auto-correlation plots indicated

white noise structure. Thus, we selected the order d∗ = 9 in the succeeding steps of our analysis.

Since no change was suspected during the pre-stimulus period (before t = 250), we used a non

symmetrical weight function and thus restricted the change point search interval to [250, 490]. The

specific weight functions gives somewhat more weight to late (post-stimulus) changes

w(t) = I[ 250512, 490512 ](t)(1− t)

−0.25, t ∈ [0, 1].

Because we are in a multiple testing situation, where we can assume independence between trials, we

apply a false discovery rate correction as proposed by [Benjamini and Hochberg, 1995] at level 5% for all

tests below.

4.3 Results

As a first step, we applied the at-most-one-change model to this EEG data. In Table 4.1, we note the

following results. The percentage of trials that suggest at least one change is over 70% which is a strong

evidence that the brain process is highly dynamic even within a one second interval (which is the length

of one trial). As one might expect, there appears to be no significant difference in this percentage between

the left and right conditions. Compared to the test statistic based on the average, the test statistic based

on the maximum produced a higher percentage of trials with change points. This difference is consistent

for both asymptotic and bootstrap-based inference.

We then plotted the estimated density of the locations of the change points for both left and right

conditions. We note in Figure 4.2 The primary peak occurs at around t = 330 for both the left and right

conditions. The secondary peak, which is of smaller magnitude, occurs at roughly t = 450. The bimodal

feature could simply be due to random variation in the change points or could suggest the presence of a

second change point in the brain process within a trial. We investigated this idea further and plotted the

densities at each of the 12 channels. We do this primarily to find out which channels best display changes

and if all channels do contain changes, we want to identify those changes. The densities obtained using

the M1-boot statistic in Figure 4.2. We observe that the estimated density plots vary across channels

- some channels are more responsive to the stimulus than others. Moreover, some channels indicate the

presence of two change points thus suggesting three regimes in the brain dynamics within a trial: the first

being the post-stimulus regime, the second is the time period somewhere between instruction and motor

response, the third is the post-motor response regime which reverts back to the first pre-stimulus regime.

Motivated by the suggestion of a second change point within a trial, we applied our proposed mul-

tivariate epidemic change point test and report the results in Table 4.2. We note that this test also

20

Page 21: Detection of Changes in Multivariate Time Series With ...ckirch/seite/publications/media/pp_kmo_1… · Detection of Changes in Multivariate Time Series With Application to EEG data

rejects and even more often than AMOC statistic which is again evidence that, for this EEG dataset,

the epidemic change model is closer to reality than the AMOC. Digging further into the behavior at the

individual channels, the plots in Figure 4.3 suggest that all channels share similar estimated densities of

change points under the epidemic model: the first change point is concentrated around t = 330 and the

second at around t = 370. Moreover, a visual inspection indicates that the first change point for the left

condition occurs before that of the right condition in some channels. Of course, this needs to be con-

firmed via formal statistical testing. There are also a number of studies on handedness and attention. For

this right-handed subject, leftward movements might be more novel and more attended than rightward

movements and this enhanced directional attention could have slightly influenced the reactivity speed.

Our collaborator considers this finding to be very interesting and certainly warrants deeper neuroscientific

investigations.

To further study the distribution of the change points under the epidemic model, we produced the

contour density plots in Figure 4.4 and noted that while some channels are able to detect changes in the

rightward movement better than others, there are also other channels that are more sensitive to changes in

the leftward movement. We also observe that the concentration of the first change point occurs at around

t = 330 and the second occurs at around 40 time points following the first, i.e., t = 370. Moreover,

for both left and right conditions, a majority of the second change points occur within 60 time points

following the first. Moreover, the plots suggest that the variation for the left conditions is slightly higher

than that for the right condition. Again, this will need to be formally tested before one can make any

strong claims. One could argue that, for right-handed individuals, leftward movements are considered to

be more novel than rightward movements which then leads to a greater variation in the brain responses.

4.4 Summary

This analysis demonstrates how our proposed procedures could be helpful in studying the complex brain

dynamics during a cognitive motor task. The results were quite interesting for our collaborator because

it raises more questions and thus motivates further experimentation. These results are also potentially

highly relevant to brain computer interface (BCI) where non-invasive signals such as EEGs serve as

inputs to a processor that controls the movement of a robotic arm. When successfully implemented, this

technology could eventually allow people with severe disabilities (such as stroke victims and war veterans

with limited movement) to control robots that can help them in daily living activities. Our work here

provides a small contribution towards understanding the dynamics of brain process during a motor task.

This inter-disciplinary collaboration is a demonstration of the importance of statistical tools in generating

more scientific hypotheses and thus pushing forward the boundaries of science.

M (1)asymp. M (2)asymp. M (1)boot. M (2)boot.

Left 0.94 0.89 0.83 0.76

Right 0.95 0.87 0.79 0.72

Table 4.1: Relative number of estimated change points under the multivariate AMOC model.

21

Page 22: Detection of Changes in Multivariate Time Series With ...ckirch/seite/publications/media/pp_kmo_1… · Detection of Changes in Multivariate Time Series With Application to EEG data

M (3)asymp. M (4)asymp. M (3)boot. M (4)boot.

Left 0.99 0.99 0.89 0.91

Right 0.97 0.95 0.92 0.91

Table 4.2: Relative number of estimated change points under the multivariate epidemic model.

5 Conclusions

Motivated by the current demands within the neuroscientific community to have a deeper understand-

ing of the changing brain dynamics during a cognitive task, we rigorously developed novel statistical

methods for detecting change points in multivariate time series. Our proposed method, inspired by

[Huskova et al., 2007], is a generalization of the class of score type change point statistics to the vector

autoregressive (VAR) case and the epidemic change alternative. Our proposed procedures do not re-

quire the observed time series to actually follow the VAR model. In fact, brain processes are unlikely

to strictly follow some VAR model. Instead, following the strategy implicitly employed by practitioners,

our approach takes model misspecification into account so that our detection procedure uses the model

background merely for feature extraction. We derive the asymptotic distributions of our test statistics and

show that our procedure has asymptotic power of 1. The proposed test statistics require the estimation of

the inverse of the (long run) auto-covariance matrix which is particularly difficult in higher-dimensional

settings which is even worse in the multivariate case because the number of entries of the positive defi-

nite weight matrix H increases polynomially with the number of dimensions. We overcome this problem

by focusing on changes that are captured by relationships of each channel Yt(i) with its own past and

by judiciously forming H to be block diagonal. Since the channels of the multivariate time series are

highly likely to be dependent, we estimate the underlying covariance structure via a regression bootstrap

method which is demonstrated here and in Huskova et al. (2008) to be stable. We investigated the finite

sample properties of our proposed method via extensive numerical experiments. Finally, we applied our

procedure to electroencephalograms and demonstrated its potential impact in identifying change points

in complex brain processes during a cognitive motor task.

250 300 350 400 450 500 550

0.000

0.005

0.010

0.015Left

M1 asymp.M2 asymp.M1 boot.M2 boot.

250 300 350 400 450 500 550

0.000

0.005

0.010

0.015Right

M1 asymp.M2 asymp.M1 boot.M2 boot.

Figure 4.1: Estimated densities of the change point under the multivariate AMOC model.

22

Page 23: Detection of Changes in Multivariate Time Series With ...ckirch/seite/publications/media/pp_kmo_1… · Detection of Changes in Multivariate Time Series With Application to EEG data

250 300 350 400 450 500 550

0.000

0.002

0.004

0.006

0.008

0.010 channel 1

250 300 350 400 450 500 550

0.000

0.002

0.004

0.006

0.008

0.010 channel 2

250 300 350 400 450 500 550

0.000

0.002

0.004

0.006

0.008

0.010 channel 3

250 300 350 400 450 500 550

0.000

0.002

0.004

0.006

0.008

0.010 channel 4

250 300 350 400 450 500 550

0.000

0.002

0.004

0.006

0.008

0.010 channel 5

250 300 350 400 450 500 550

0.000

0.002

0.004

0.006

0.008

0.010 channel 6

250 300 350 400 450 500 550

0.000

0.002

0.004

0.006

0.008

0.010 channel 7

250 300 350 400 450 500 550

0.000

0.002

0.004

0.006

0.008

0.010 channel 8

250 300 350 400 450 500 550

0.000

0.002

0.004

0.006

0.008

0.010 channel 9

250 300 350 400 450 500 550

0.000

0.002

0.004

0.006

0.008

0.010 channel 10

250 300 350 400 450 500 550

0.000

0.002

0.004

0.006

0.008

0.010 channel 11

250 300 350 400 450 500 550

0.000

0.002

0.004

0.006

0.008

0.010 channel 12

LeftRight

Figure 4.2: Weighted density estimates of the change point for each channel under the AMOC model using the

M1- bootstrap procedure.

References

[Aston and Kirch, 2013] Aston, J. and Kirch, C. (2013). Change points in high-dimensional settings.

Diskussions-Papier, KIT.

[Aue et al., 2009] Aue, A., Hormann, S., Horvath, L., and Reimherr, M. (2009). Break detection in the

covariance structure of multivariate time series model. Ann. Statist., 37:4046–4087.

[Bai and Perron, 1998] Bai, J. and Perron, L. (1998). Estimating and testing linear models with multiple

structural changes. Econometrica, 66:47–78.

[Benjamini and Hochberg, 1995] Benjamini, Y. and Hochberg, J. (1995). Controlling the false discovery

rate: A practical and powerful approach to multiple testing. J. Royal Stat. Soc., Ser. B, 57:289–300.

[Bickel and Buhlman, 1999] Bickel, P. and Buhlman, P. (1999). A new mixing notion and functional

central limit theorems for a sieve bootstrap in time series. Bernoulli, 5:413–446.

[Brockwell and Davis, 1991] Brockwell, P. and Davis, R. (1991). Time Series: Theory and Methods.

Springer, New York, second edition.

[Csorgo and Horvath, 1997] Csorgo, S. and Horvath, L. (1997). Limit Theorems in Change-point Analy-

sis. Wiley, Chichester.

[Davis et al., 1995] Davis, R., Huang, D., and Yao, Y. (1995). Testing for a change in the parameter

values and order of an autoregressive model. Ann. Statist., 23:282–304.

23

Page 24: Detection of Changes in Multivariate Time Series With ...ckirch/seite/publications/media/pp_kmo_1… · Detection of Changes in Multivariate Time Series With Application to EEG data

250 300 350 400 450 500 550

0.000

0.002

0.004

0.006

0.008

0.010

0.012

0.014

channel 1

250 300 350 400 450 500 550

0.000

0.002

0.004

0.006

0.008

0.010

0.012

0.014

channel 2

250 300 350 400 450 500 550

0.000

0.002

0.004

0.006

0.008

0.010

0.012

0.014

channel 3

250 300 350 400 450 500 550

0.000

0.002

0.004

0.006

0.008

0.010

0.012

0.014

channel 4

250 300 350 400 450 500 550

0.000

0.002

0.004

0.006

0.008

0.010

0.012

0.014

channel 5

250 300 350 400 450 500 550

0.000

0.002

0.004

0.006

0.008

0.010

0.012

0.014

channel 6

250 300 350 400 450 500 550

0.000

0.002

0.004

0.006

0.008

0.010

0.012

0.014

channel 7

250 300 350 400 450 500 550

0.000

0.002

0.004

0.006

0.008

0.010

0.012

0.014

channel 8

250 300 350 400 450 500 550

0.000

0.002

0.004

0.006

0.008

0.010

0.012

0.014

channel 9

250 300 350 400 450 500 550

0.000

0.002

0.004

0.006

0.008

0.010

0.012

0.014

channel10

250 300 350 400 450 500 550

0.000

0.002

0.004

0.006

0.008

0.010

0.012

0.014

channel11

250 300 350 400 450 500 550

0.000

0.002

0.004

0.006

0.008

0.010

0.012

0.014

channel12

1. CP2. CP

LeftRight

Figure 4.3: Estimated density, weighted to integrate to the relative number of rejections, of the first and second

change point under the epidemic model using the T1-bootstrap procedure.

[Davis et al., 2006] Davis, R., Lee, T., and Rodriguez-Yam, G. (2006). Structural break estimation for

nonstationary time series models. J. Amer. Stat. Assoc., 101:223–229.

[Fan and Yao, 2003] Fan, J. and Yao, Q. (2003). Nonlinear Time Series: Nonparametric and Parametric

Methods. Springer, New York.

[Horvath, 1993] Horvath, L. (1993). Change in autoregressive processes. Stoch. Proc. Appl., 5:221–242.

[Horvath, L. et al., 1999] Horvath, L., Kokoszka, P., and Steinebach, J. (1999). Testing for changes in

multivariate dependent observations with an application to temperature changes. J. Multivariate Anal.,

68:96–119.

[Huskova et al., 2007] Huskova, M., Praskova, Z., and Steinebach, J. (2007). On the detection of changes

in autoregressive time series, I. Asymptotics. J. Statist. Plann. Infer., 137:1243–1259.

[Huskova and Slaby, 2001] Huskova, M. and Slaby, A. (2001). Permutation test for multiple changes.

Kybernetika, 37:606–622.

[Kirch and Muhsal, 2011] Kirch, C. and Muhsal, B. (2011). A MOSUM procedure for the estimation of

multiple deterministic as well as random change points. Diskussions-Papier, KIT.

[Kuelbs and Philipp, 1980] Kuelbs, J. and Philipp, W. (1980). Almost sure invariance principles for

partial sums of mixing b-valued random variables. Ann. Probab., 8:1003–1036.

24

Page 25: Detection of Changes in Multivariate Time Series With ...ckirch/seite/publications/media/pp_kmo_1… · Detection of Changes in Multivariate Time Series With Application to EEG data

250 300 350 400 450

0

50

100

150

200

250

25

50

75 95

●●

●●

●●●

●●

●●

●● ●●

● ●

●●

●●

●●

● ●

● ●●

●●

●●

Left

250 300 350 400 450

0

50

100

150

200

250

25

50

50

50

75

95

●●

●●●

●●

●●

● ●●

●●●

●●

●●

● ●

●●

● ●

● ●●

●●

● ●● ●

●●

●●

●●

●●

●●

●●

●●

Right

Figure 4.4: Contour plot of the joint density of 1.CP and 2.CP-1.CP

[Kulperger, 1985] Kulperger, R. (1985). On the residuals of autoregressive processes and polynomial

regression. Stoch. Process. Appl., 21:107–118.

[Lesage et al., 2002] Lesage, P., Glangeaud, F., and Mars, J. (2002). Applications of autoregressive models

and time-frequency analysis to the study of volcanic tremor and long-period events. J. Volcan. Geo.

Res., 115:391–417.

[Marconi et al., 2001] Marconi, B., Genevesio, A., Battaglia-Meyer, A., Ferraina, S., and Caminiti, R.

(2001). Eye-hand coordination during reacning: Anatomical relationships between the parietal and

frontol cortex. Cerebral Cortex, 11:513–527.

[Marusiakova, 2009] Marusiakova, M. (2009). Ph.d. thesis: Tests for multiple changes in linear regression

models. Charles University in Prague, Faculty of Mathematics and Physics.

[Pfurtscheller and Haring, 1972] Pfurtscheller, G. and Haring, G. (1972). The use of an eeg autoregressive

model for the time-saving calculation of spectral power density distributions with a digital computer.

Elec. Clin. Neurophys., 33:113–115.

[Politis, 2003] Politis, D. (2003). Adaptive bandwidth choice. J. Nonpar. Stat., 15:517–533.

[Preußs et al., 2013] Preußs, P., Puchstein, R., and Dette, H. (2013). Detection of multiple structural

breaks in multivariate time series. Ruhr-Universitat Bochum, Fakultat fur Mathematik.

[Vermaak et al., 2002] Vermaak, J., Andrieu, C., Doucet, A., and Godsill, S. (2002). Particle methods for

bayesian modeling and enhancement of speech signals. IEEE Trans. Speech Audio Proc., 10:173–185.

6 Appendix

6.1 Proofs of Section 2.1.4

Lemma 6.1. Let Assumption A.1 hold. Under the alternative additionally assume A.2 and A.5.

25

Page 26: Detection of Changes in Multivariate Time Series With ...ckirch/seite/publications/media/pp_kmo_1… · Detection of Changes in Multivariate Time Series With Application to EEG data

a) Under the null hypothesis as well as both alternatives, it holds

an(l)→ a(l) a.s., l = 1, . . . , d,

where under the null hypothesis a(l) = a1(l).

b) Under the null hypothesis we additionally get

an(l)− a(l) = OP (n−1/2), l = 1, . . . , d.

Remark 6.1. If the time series after the change point fulfills also assumptions as in A.1, then assertion

b) also remains true under both alternatives.

Proof. By (2.6) it holds an(l) = C−1n (l) 1n

∑ni=1Xi−1(l)Yi(l). Since mixing sequences are ergodic by

the ergodic theorem (see [Fan and Yao, 2003], Prop. 2.8)

Cn(l)→ Q(l) a.s., qn(l) :=1

n

n∑i=1

Xi−1(l)Yi(l)→ q(l) a.s.,

which proves a) since a(i) = Q−1(i)q(i) under both the null hypothesis as well a alternatives. For b)

we note that Assumptions A.1 imply an invariance principle (see [Kuelbs and Philipp, 1980] Theorem 4)

which in turn implies a central limit theorem showing that

√n(qn(l)− c1(l)) = OP (1),

√n(Cn(l)−C1(l)) = OP (1).

This implies

C−1n (l) qn(l)−C−11 (l)c1(l)

= C−1n (l)(C1(l)− Cn(l))C−11 (l)qn(l) + C−11 (qn(l)− c1(l)) = OP

(n−1/2

),

which concludes the proof.

Lemma 6.2. Let the null hypothesis and Assumption A.1 hold.Then:

a) The following functional central limit theorem holds: 1√n

bnsc∑i=1

ξi, s ∈ [0, 1]

D[0,1]−→ {B(s) : 0 6 s 6 1},

ξt = (ξT

t (1), . . . , ξT

t (d))T = ((Xt−1(1))T et(1), . . . , (Xt−1(d))T et(d))T ,

where B(·) is a dp-dimensional Brownian bridge with covariance matrix Σ.

b) It holds for all 0 6 β < 12

max16k6n

n2β−1/2

kβ(n− k)β

∥∥∥∥∥k∑i=1

ξi

∥∥∥∥∥ = OP (1).

26

Page 27: Detection of Changes in Multivariate Time Series With ...ckirch/seite/publications/media/pp_kmo_1… · Detection of Changes in Multivariate Time Series With Application to EEG data

Proof. First, note that {ξt : t} as well as {Xt(l)XTt (l) : t} are by Assumption A.1 strong mixing with

the same mixing rate and 2 + ν/2 existing moments. Consequently, the assumptions of the invariance

principle in [Kuelbs and Philipp, 1980], Theorem 4, are fulfilled. This implies the functional central limit

theorem 1√n

bntc∑i=1

ξt, t ∈ [0, 1]

D[0,1]−→ {W(t) : 0 6 t 6 1}, (6.1)

where W (·) is a dp-dimensional Wiener process with covariance matrix Σ. Furthermore, in connection

with a standard Hajek-Renyi-inequality it implies for all 0 6 β < 1/2

max16k6n

n2β−1/2

kβ(n− k)β

∣∣∣∣∣k∑i=1

(ζi −

1

n

n∑t=1

ζt

)∣∣∣∣∣ = OP (1), (6.2)

for ζt any channel of Xt−1(l)et(l), l = 1, . . . , p, as well as any channel of Xt−1(l)XTt−1(l), l = 1, . . . , p. As the

minimizer of the least squares equations an(l) solve the corresponding score equations, i.e.∑n

i=1 ξi(l) = 0,

l = 1, . . . , d. Consequently,

k∑i=1

ξi(l) =

k∑i=1

ξi(l)− 1

n

n∑j=1

ξj(l)

=

k∑i=1

ξi(l)− 1

n

n∑j=1

ξj(l)

− (an(l)− a1(l))T

k∑i=1

Xi(l)XTi (l)− 1

n

n∑j=1

Xj(l)XTj (l)

.

Now, Assertion a) follows by (6.1) and (6.2) with β = 0 together with Lemma 6.1 a). Assertion b) follows

by (6.2) and Lemma 6.1 a).

Proof of Theorem 2.1. We will only sketch the proof for the case where w(·) has exactly one

discontinuity point 0 < u < 1. Define wl(t) = w(t−) and wr(t) = w(t+), then

sup06t6u

|wl(bntc/n)− wl(t)| → 0, supt>dnue/n

|wr(bntc/n)− wr(t)| → 0. (6.3)

For any 0 < τ < min(u, 1 − u) it holds supτ6t61−τ w2(t) < ∞, hence by Lemma 6.2 a), H

P−→ H and

(6.3) (for w(·) right continuous in u - otherwise allow for equality in the second supremum)

maxbτnc6k6n−bτnc

w2(k/n)

nZTk HZk

= max

(supτ6t6u

w2(bntc/n)

nZTbntcHZbntc, sup

dnue/n<t6n−bτnc

w2(bntc/n)

nZTbntcHZbntc

)

D−→ supτ6t61−τ

w2(t)

pd∑j=1

B2j (t),

where we use the a.s. continuity of a Brownian bridge. By Lemma 6.2 b) we get for 1/2 > β > α (as in

Assumption A.3)

max16k6bτnc

w2(k/n)

nZTk HZk = OP (1)

(sup06t6τ

tβw(t)

)2

,

27

Page 28: Detection of Changes in Multivariate Time Series With ...ckirch/seite/publications/media/pp_kmo_1… · Detection of Changes in Multivariate Time Series With Application to EEG data

where the rates are uniform in τ . Assumption A.3 implies for β > α

sup06t6τ

tβw(t)→ 0 as τ → 0.

Similar assertions can be obtained for k > n−bτnc as well as for the corresponding expressions involving

the Brownian bridges in the limit. Standard arguments then conclude the proof of a). Assertion b) can

be dealt with analogously.

Proof of Theorem 2.2. This follows immediately from Lemma 6.2 a).

6.2 Proofs of Section 2.2

The key to understanding the behavior under alternatives is the following lemma:

Lemma 6.3. Let Assumptions A.1, A.2 and A.5 be fulfilled. Let λ be as in (2.12) and the alternative

hold.

a) Then, we get

c1(l)−C1(l)a(l) = −1− λλ

(c2(l)−C2(l)a(l)) .

b) Under the AMOC alternative we get

sup06t61

∥∥∥∥∥∥ 1

n

btnc∑j=1

ξj(l)− gA(t) (c1(l)−C1(l)a(l))

∥∥∥∥∥∥2

= op(1),

where

gA(t) =1

1− λ

t (1− λ), t 6 λ,

λ (1− t), t > λ.

c) Under the epidemic change alternative we get

sup06t1<t261

∥∥∥∥∥∥ 1

n

bt2nc∑j=bt1nc

ξj(l)− gB(t1, t2) (c1(l)−C1(l)a(l))

∥∥∥∥∥∥2

= op(1),

where gB(t1, t2) = gB(t2)− gB(t1) and

gB(t) =1

λ2 − λ1

t (λ2 − λ1), t 6 λ1,

λ1 − t (1− λ2 + λ1), λ1 6 t 6 λ2,

(t− 1) (λ2 − λ1), t > λ2.

Proof. Assertion a) is equivalent to q(l) − Q(l)a(l) = 0 which holds by definition of a(l). By

Lemma 6.1 a)

1

n

btnc∑j=1

ξj(l) =1

n

bntc∑j=1

Xj−1(l)(Yj(l)− XTj−1(l)a(l))− (an(l)− a(l))1

n

btnc∑j=1

Xj−1(l)XTj−1(l)

=1

n

btnc∑j=1

Xj−1(l)Yj(l)−

1

n

btnc∑j=1

Xj−1(l)XTj−1(l)

a(l) + oP (1),

28

Page 29: Detection of Changes in Multivariate Time Series With ...ckirch/seite/publications/media/pp_kmo_1… · Detection of Changes in Multivariate Time Series With Application to EEG data

where the latter rate is uniform in t. By the ergodic theorem we get

sup06t6λ

∣∣∣∣∣∣ 1nbtnc∑j=1

Xj−1(l)Yj(l)− c1(l)t

∣∣∣∣∣∣→ 0 a.s.

as well as

supλ6t61

∣∣∣∣∣∣ 1nbλnc∑j=1

Xj−1Yj +

btnc∑j=bλnc

Xj−1Yj

− (λc1(l) + (t− λ)c2(l))

∣∣∣∣∣∣→ 0 a.s.

A similar assertion holds for the second sum showing that

sup06t61

∣∣∣∣∣∣ 1nbtnc∑j=1

ξj(l)−max(t, λ)(c1(l)−C1(l)a(l))− (t− λ)+(c2(l)−C2(l)a(l))

∣∣∣∣∣∣→ 0 a.s.,

which gives assertion b) using the formula in a). Assertion c) can be obtained analogously.

Proof of Theorem 2.3. Let 0 < t0 < 1 a continuity point with w(t0) > 0, then we get by Lemma 6.3

M (1)n > n

(∆2Hw2(t0)g

2A(t0) + oP (1)

)P−→∞,

M (2)n = n

(∆2H

∫ 1

0w2(t)g2A(t) dt+ oP (1)

)P−→∞,

where for the second statement, we split the sum at possible discontinuity points and use (6.3). The

assertions for b) follow analogously.

Proof of Theorem 2.4. Define

∆2 = (c1 − diag(C1(1), . . . ,C1(d))a)T H−1A (c1 − diag(C1(1), . . . ,C1(d))a) .

We sketch the proof of a) for exactly one (left-continuous – otherwise allow for equality in the second

supremum) discontinuity point u. By sup06t61w2(t) <∞ and (6.3) we get by Lemma 6.3

sup06t6u,dnue/n<t61

∣∣∣∣w2 (bntc/n)

n2ZTbntcH

−1Zbntc −∆2w2(t)g2A(t)

∣∣∣∣ = oP (1). (6.4)

Since by assumption w2(t)g2A(t) has a unique supremum at λ, which will eventually be in {0 6 t 6

u, dnue/n < t 6 1}, and the estimator is by definition in that set, standard arguments can be adapted

to give assertion a). For b) note that gB(t) is continuous and has a unique maximum at t = λ1 and a

unique minimum at t = λ2, hence gB(t1, t2) = gB(t2)−gB(t1) is continuous and has a unique (for t1 < t2)

maximum at (λ1, λ2). From this assertion b) follows by standard arguments.

Proof of Remark 2.2. Similarly as in Lemma 6.2, we obtain for 0 6 β < 1/2

max16k6n

n2β+1/2

kβ(n− k)β

∥∥∥∥∥∥ 1

n

k∑j=1

ξj(l)− gA(k/n) (c1(l)−C1(l)a(l))

∥∥∥∥∥∥ = OP (1)

from which we get (6.4), so that we can conclude as in the proof of Theorem 2.4.

29


Recommended